ArticlePDF Available

Exhaust Waste Heat Recovery from a Heavy-Duty Truck Engine: Experiments and Simulations

Authors:
  • TitanX Engine Cooling AB

Abstract and Figures

Waste heat recovery using an (organic) Rankine cycle is an important and promising technology for improving engine efficiency and thereby reducing the CO2 emissions due to heavy-duty transport. Experiments were performed using a Rankine cycle with water for waste heat recovery from the exhaust gases of a heavy-duty Diesel engine. The experimental results were used to calibrate and validate steady-state models of the main components in the cycle: the pump, pump bypass valve, evaporator, expander, and condenser. Simulations were performed to evaluate the cycle performance over a wide range of engine operating conditions using three working fluids: water, cyclopentane, and ethanol. Additionally, cycle simulations were performed for these working fluids over a typical long haul truck driving cycle. The predicted net power output with water as the working fluid varied between 0.5 and 5.7 kW, where the optimal expander speed was dependent on the engine operating point. The net power output for simulations with cyclopentane was between 1.8 and 9.6 kW and that for ethanol was between 1.0 and 7.8 kW. Over the driving cycle, the total recovered energy was 11.2, 8.2, and 5.2 MJ for cyclopentane, ethanol, and water, respectively. These values correspond to energy recoveries of 3.4, 2.5, and 1.6 %, respectively, relative to the total energy requirement of the engine. The main contribution of this paper is the presentation of experimental data on a complete Rankine cycle-based WHR system coupled to a heavy-duty engine. These results were used to validate component models for simulations, allowing for a realistic estimation of the steady-state performance under a wide range of operating conditions for this type of system.
Content may be subject to copyright.
Exhaust waste heat recovery from a heavy-duty truck engine:
Experiments and simulations
Jelmer Rijpkema
a
,
*
, Olof Erlandsson
b
, Sven B. Andersson
a
, Karin Munch
a
a
Mechanics and Maritime Sciences, Chalmers University of Technology, Gothenburg, Sweden
b
TitanX Engine Cooling AB, S
olvesborg, Sweden
article info
Article history:
Received 4 March 2021
Received in revised form
22 July 2021
Accepted 4 August 2021
Available online 14 August 2021
Keywords:
Experiments
Exhaust gases
Heavy-duty
Internal combustion engine
Organic rankine cycle (ORC)
Reciprocating piston expander
Long haul truck
Waste heat recovery
abstract
Waste heat recovery using an (organic) Rankine cycle is an important and promising technology for
improving engine efciency and thereby reducing the CO
2
emissions due to heavy-duty transport. Ex-
periments were performed using a Rankine cycle with water for waste heat recovery from the exhaust
gases of a heavy-duty Diesel engine. The experimental results were used to calibrate and validate steady-
state models of the main components in the cycle: the pump, pump bypass valve, evaporator, expander,
and condenser. Simulations were performed to evaluate the cycle performance over a wide range of
engine operating conditions using three working uids: water, cyclopentane, and ethanol. Additionally,
cycle simulations were performed for these working uids over a typical long haul truck driving cycle.
The predicted net power output with water as the working uid varied between 0.5 and 5.7 kW, where
the optimal expander speed was dependent on the engine operating point. The net power output for
simulations with cyclopentane was between 1.8 and 9.6 kW and that for ethanol was between 1.0 and
7.8 kW. Over the driving cycle, the total recovered energy was 11.2, 8.2, and 5.2 MJ for cyclopentane,
ethanol, and water, respectively. These values correspond to energy recoveries of 3.4, 2.5, and 1.6%,
respectively, relative tothe total energy requirement of the engine. The main contribution of this paper is
the presentation of experimental data on a complete Rankine cycle-based WHR system coupled to a
heavy-duty engine. These results were used to validate component models for simulations, allowing for a
realistic estimation of the steady-state performance under a wide range of operating conditions for this
type of system.
©2021 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).
1. Introduction
The continuous increase of anthropogenic greenhouse gas
(GHG) emissions is causing severe adverse effects on the climate.
Consequently, the global energy system must rapidly reduce its
emissions [1]. Because heavy-duty (HD) trucks and buses are
responsible for over 5% of the total GHG emissions in Europe [2],
ways to reduce emissions from HD vehicles are needed. As a result,
the European Union has imposed CO
2
emission standards for HD
vehicles that require emission reductions of 15% from 2025 on-
wards and 30% from 2030 onwards, relative to the 2019 baseline
[3]. Several technologies and powertrain concepts have been pro-
posed to help meet these requirements, including improvements in
combustion and air management efciency, predictive powertrain
control, hybridization, reduction of friction and other losses,
renewable fuels, hydrogen, fuel cells, and waste heat recovery
(WHR) [4,5]. WHR systems generate power from the waste heat of
the heat sources in a HD engine, namely the charge air cooler (CAC),
the exhaust recirculation (EGR) cooler, the engine coolant, or the
exhaust gases [6]. Many different WHR concepts and technologies
exist, including turbocompounding [7], thermoacoustic convertors
[8], thermoelectric generators [9], and technologies based on
thermodynamic cycles such as the Brayton [10], Stirling [11],
Rankine [12], or various ash cycles [6]. Systems based on the
(organic) Rankine cycle (ORC) have been found to achieve good
performance and exibility, although the added weight,
complexity, and payback time remain obstacles to their commercial
implementation [13].
In an ORC-based WHR system, waste heat is used to evaporate a
working uid at elevated pressure. The high pressure, high tem-
perature uid is then expanded, converting the heat energy into
power, after which the uid is condensed before entering the
*Corresponding author.
E-mail address: jelmer.rijpkema@chalmers.se (J. Rijpkema).
Contents lists available at ScienceDirect
Energy
journal homepage: www.elsevier.com/locate/energy
https://doi.org/10.1016/j.energy.2021.121698
0360-5442/©2021 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
Energy 238 (2022) 121698
pump. The goal of the WHR system is to maximize the net power
output, while maintaining superheated vapor conditions at the
inlet of the expander [14]. The performance of such systems in
heavy-duty Diesel (HDD) engines has been the topic of many
publications (mostly simulation studies), with reported fuel savings
of over 5% [13,15]. Important issues addressed in these studies
include heat source selection [16], the choice of working uid
[16e18], heat exchanger modeling [19,20], expander selection and
performance [21], cycle conguration [22], and techno-economic
performance optimization [23]. Due to the transient operation of
an internal combustion engine, much recent research has focused
on developing and validating dynamic models and controls for ORC
systems in HD engines [14,24e27]. In addition to the academic
interest, ORC systems for HDD engines have attracted considerable
interest from major automotive industry rms in the last decade,
including AVL [28], BMW [29,30], Bosch [31], Cummins [32,33],
Daimler [34], Honda [35], Mahle [36], Scania [37], Volkswagen [38],
Volvo Car Corporation [39,40], and Volvo Group [24,41].
A number of notable recent publications concerning experi-
mental research on WHR from HDD engines are briey discussed
below and listed in Table 1. Seher et al. [31] reported an experi-
mental study on a Rankine cycle with water for a 12 L HDD engine.
A maximum power output of 14 kW was obtained with a piston
expander, corresponding to 4.3% of the engine power, while 9 kW
(2.8%) was achieved with a turbine expander. Based on simulations
it was concluded that water or ethanol with a piston or ethanol
with a turbine were the preferred solutions, giving a maximum
relative power of 5.3%. Furukawa et al. [42] tested two ORC systems
for WHR from a downsized HDD engine. In their rst system, heat
was recovered from the engine coolant and EGR gases, reducing
fuel consumption by 3.8%. In their second system, the coolant
temperature was increased from 86 to 105
C, the exhaust gas was
included as a heat source, and a recuperator was added, improving
the fuel consumption reduction to 7.5%. Yang et al. [43] used an ORC
with R245fa and a screw expander to recover up to 28.6 kW (10.2%)
from a HDD engine. Zhang et al. [44] also used a screw expander in
an ORC with R123 for exhaust heat recovery from a HDD engine and
achieved a maximum power of 10.4 kW. Bettoja et al. [41] per-
formed experiments on two systems for two different engines: a
Volvo US10 and a CRF Cursor 11. For the Volvo engine, heat was
recovered from the exhaust and EGR with a water/ethanol mixture.
An orice was used instead of an expander and the system was
predicted to achieve a relative power recovery between 1.5 and 3%.
For the CRF engine, a system with R245fa and a turbine was used,
giving a maximum power output of 2.5 kW. Latz et al. [45] used a
similar setup as in this paper featuring the same engine, but
recovered heat from the EGR cooler using a reciprocating piston
expander and water. The maximum recovered power in this case
was 2.7 kW. Simulations were performed to identify important
parameters for performance improvement of the EGR evaporator
and piston expander. Shu et al. [46] compared the working uids
R123 and R245fa in a WHR system with an intermediate oil loop
connected to the exhaust of an HDD engine. No expander was
installed in the system, but, by using estimated efciencies, a
maximum power output of 9.7 kW for R123 was predicted. In
another study, Yu et al. [47] replaced the oil loop with another
Rankine cycle with water. This improved the estimated power
output to 12.7 kW or 5.6% in relative terms. More recently, Shi et al.
[48] used the same engine for WHR from the exhaust and engine
coolant with four congurations of a CO
2
transcritical Rankine cy-
cle. Using estimated efciencies, the maximum power output was
predicted to be 3.5 kW. Guillaume et al. [49] simulated the exhaust
conditions of a HDD engine using a boiler with thermal oil and
concluded that R1233zd(E) performed better than R245fa in their
experiments, providing a maximum power output of 2.8 kW with a
turbine expander. Alshammari et al. [50] recovered the exhaust
heat using an ORC with R1233zd(E), a turbine, and an intermediate
thermal oil loop, giving a maximum recovered power of 6.3 kW.
Their results were complemented by CFD simulations and evalua-
tions of the radial inow turbine performance. In a more recent
study [51], the same group subsequently tested the same engine
with a WHR system featuring a thermal loop, a turbine, and
Novec649 as the working uid. The maximum power output in this
case was 9.1 kW (11.2%). Additionally, they showed that increasing
the cooling water temperature and superheating temperature
reduced the performance of the turbine.
Although there have been many publications regarding the use
of ORC for WHR in HDD engines, there is still a lack of publicly
available experimental data. This is partly because publications
from industry often report performance improvements without
providing much detail on the cycle components. Additionally, many
publications reporting results of dynamic models often present
experimental data that was used to validate the model. These re-
sults only specify the controlled parameters (e.g. mass ow rate or
evaporator outlet temperature) without offering insight into the
cycle or component performance. In this paper, an experimental
Table 1
Recent experimental studies on WHR from HDD engines using ORCs.
Reference Year Engine Heat source Fluids(s) Expander _
W
max
_
W
max
_
W
eng
- - L/kW - - - kW %
Seher et al. [31] 2012 12.0/326 Exhaust Water Piston, Turbine 14 4.3
Furukawa et al. [42] 2014 9.0/- Exhaust, EGR, HFE Turbine e7.5
Coolant
Yang et al. [43] 2014 9.7/280 Exhaust R245fa Screw 28.6 10.2
Zhang et al. [44] 2014 -/250 Exhaust R123 Screw 10.4 4.2
Bettoja et al. [41] 2016 11.1/353 Exhaust R245fa Turbine 2.5 e
12.7/317 Exhaust, EGR Water/Ethanol ee3.0
Latz et al. [45] 2016 12.8/373 EGR Water Piston 2.7
Shu et al. [52] 2016 8.4/243 Exhaust R123, R245fa e9.7 4.0
Yu et al. [47] 2016 8.4/243 Exhaust Water, R123 e12.7 5.6
Guillaume et al. [49] 2017 -/ - Exhaust R245fa, Turbine 2.8 e
R1233zd(E)
Shi et al. [48] 2017 8.4/243 Exhaust, CO
2
e3.5 e
Coolant
Alshammari et al. [50] 2018 7.3/206 Exhaust R1233zd(E) Turbine 6.3 7.6
Alshammari et al. [51] 2019 7.3/206 Exhaust Novec649 Turbine 9.1 11.2
J. Rijpkema, O. Erlandsson, S.B. Andersson et al. Energy 238 (2022) 121698
2
setup consisting of a Rankine cycle with water connected to the
exhaust of a HDD engine is evaluated and used to calibrate and
validate models of the relevant cycle components. The components
are then used to develop a realistic Rankine cycle model to deter-
mine the optimum performance in a driving cycle for a variety of
working uids. Simulation of the driving cycle allows for an eval-
uation of the performance of the WHR system over the full oper-
ational range of the engine. However, the corresponding transient
effects are considered outside the scope of this paper, and are,
therefore, not taken into account. The goal of this publication is to
provide detailed experimental data on a full Rankine system con-
nected to a HDD engine, and to develop models that give an ac-
curate prediction of the WHR system performance under realistic
operating conditions.
2. Experimental setup
The experimental setup is shown in Fig. 1 and a schematic
overview is presented in Fig. 2. The setup consists of a heavy-duty
truck engine whose exhaust gases are used as the heat input for a
WHR system based on a Rankine cycle using water as the working
uid. Both the engine and WHR system are placed in an engine test
cell in which the temperature and pressure can be regulated. Only
steady-state measurements are available from the engine due to
limitations on the engine brake. The setup is monitored and
controlled from the adjacent control room using multiple modules
installed in two National Instruments CompactRIO 9074 controllers
coupled to a Labview interface. Several cameras and a connecting
window allow for visual observations while running the setup.
Sensor data was measured at a sampling frequency of 10 Hz which
was written to disk every second. For each measuring point, three
minutes of data were collected and averaged.
The engine is a turbocharged 12.8 L Volvo Diesel engine with
charge air cooling (CAC) and exhaust gas recirculation (EGR); its
specications are shown in Table 2. A Schenck D900-1e water brake
is used to control the engine speed. The engine torque is controlled
by regulating the fuel ow through manual operation of the gas
pedal. Fuel is provided from a Diesel tank located in a separate fuel
storage.
Fig. 2 shows the main components of the WHR system; their
specications are listed in Table 3 together with the corresponding
controller where applicable. The WHR system is a typical Rankine
cycle with water as the working uid. The suction side of the pump
is connected to the buffer tank, which is open to the atmosphere. A
controllable pump bypass valve (BPV) is installed because the ow
at the minimum pump speed would otherwise be too large to
permit full evaporation under low load engine operating condi-
tions. The evaporator uses the exhaust gases downstream the en-
gine turbocharger to evaporate and superheat the water from the
pump. It is specically manufactured for this experimental setup by
TitanX and is shown in Fig. 3. Under start-up and shut-down con-
ditions, the uid leaving the evaporator is not always fully evapo-
rated. Therefore, the expander inlet and outlet valves are closed and
the controllable expander bypass valve (BPV) open so the ow
bypasses the expander, preventing liquid from entering. The
expander is a reciprocating piston type with two cylinders using a
separate crankcase with its own oil circuit. Water enters the
expander as superheated steam and exits as a two-phase mixture at
low pressure. During expansion, the expander converts some of the
energy in the steam into power via an electric motor, which con-
trols the expander speed. Since some of the steam enters the
crankcase of the expander, the oil is heated to 140
C, causing this
water to evaporate and be expelled to the environment. To prevent
hot oil from entering the oil pump, the oil is subsequently cooled. In
the cycle, the oil is separated from the water and the low-pressure
two-phase mixture enters the condenser, where it is condensed
and subcooled to around 15
C using process water from the test
cell. From the condenser, the subcooled water enters the buffer
tank. To avoid overpressures in the system, safety valves (SV) are
installed on the high and low pressure sides of the cycle.
Fig. 2 also shows the locations of the different sensors in the
system; the details and accuracies are shown in Table 4. The engine
speed and torque were taken from the Schenck D900-1e mea-
surements and the fuel ow using an AVL 730 fuel balance. The
engine inlet air ow was measured using the pressure drop over a
calibrated venturi tube, sufciently upstream the turbocharger to
avoid ow pulsations. Pressures in the system were measured using
WIKA A-10 pressure transmitters with different ranges depending
Fig. 1. Experimental setup.
Fig. 2. Schematic depiction of the experimental setup.
Table 2
Engine specications.
Type Volvo D13 US 2010
Conguration 4 Stroke/6 Cylinder inline/EGR
Peak power 373 kW (500 hp)
Peak torque 2373 Nm
Compression ratio 16.0:1
Bore x Stroke 131 158 mm
Displacement 12.8 L
Aspiration Turbocharged
J. Rijpkema, O. Erlandsson, S.B. Andersson et al. Energy 238 (2022) 121698
3
on the location. Temperature measurements were taken with 3 mm
diameter RS Pro Type K thermocouples. In the Rankine cycle, the
mass ow was measured with a coriolis mass ow meter and the
expander speed and torque were both measured with an universal
digital torque transducer.
3. Experimental results and component calibration
In this section the experimental results and modeling relations
of the main components in the WHR system are combined in
separate sections: Engine,Pump,Pump Bypass Valve,Evaporator,
Expander, and Condenser. If the results are used to calibrate and/or
validate the component, it is also shown. The last section shows the
comparison of the experiments and simulations for the full cycle
performance. A total of six different experimental sets were used, as
listed and described in Table 5, which shows the corresponding
number of experimental points and a description for each experi-
mental set.
Multiple quantities derived from the performed measurements
are presented in the subsequent sections. The standard deviations
for the measured (i.e. non-derived) data are represented by error
bars in the gures. Table 6 shows the maximum measurement error
based on the standard rules for error propagation [53].
3.1. Engine
The experimental measurements of the exhaust mass ow and
outlet temperature are shown in Figs. 4 and 5. These results are
averages based on the measurements acquired in experimental set
5 (see Table 5). The engine operating points are named in accor-
dance with the conventions of the European Stationary Cycle (ESC):
the letters A, B, and C indicate different engine speeds, and the
numbers 25, 50, 75, and 100 indicate the load percentages at the
corresponding speed [54]. An additional highway (HW) point was
tested, which represents typical engine conditions during highway
driving. These measurements were used to dene the heat input
conditions used in the cycle model. The original measured mass
ows were somewhat higher than those obtained in previous
experimental studies on similar engines, suggesting that the mass
ow values measured in the test cell were systematically biased
upward. Consequently, the original mass ow values were multi-
plied by an error factor of 0.75 to obtain more realistic values. The
mass ows presented in Fig. 4 have been corrected in this manner.
3.2. Pump
The mass ow from the pump is determined from its inlet
density (
r
in
) and volume ow ( _
V
pmp
) using Eq. (1). The specica-
tions of the pump are shown in Table 7.
_
m
pmp
¼
r
in
_
V
pmp
(1)
Although the axial piston pump is relatively insensitive to
pressure changes, the actual volume ow ( _
V
pmp
) is calculated by
applying a correction to the theoretical ow ( _
V
th
):
_
V
pmp
¼
_
V
th
_
V
corr
p
pmp;out
p
max
(2)
The theoretical ow in L/min is dened as:
_
V
th
¼V
pmp
N
pmp
(3)
The ow correction in L/min depends on the pump outlet
pressure (p
pmp, out
) in bar and the pump speed (N
pmp
) in rpm. This is
Table 3
Specications of cycle components.
Component Brand Type Controller
Condenser Modine Plate, counter-current ow
Evaporator TitanX Plate, cross-counter ow
Expander Voith Reciprocating piston, 2-cylinder
Expander bypass valve Swagelok SS-18RS8 Integral-bonnet needle Hanbay MCL-000 AF
Expander electric motor David McClure LTD 400 V, 3-phase, 37 kW Parker DC590þIntegrator 2
Pump Danfoss PAH2 Axial piston
Pump bypass valve Swagelok SS-1RS4 Integral-bonnet needle Hanbay MCL-000 AF
Pump electric motor Hoyer HMA2 90L-4 230 V, 3-phase, 1.5 kW IMO iDrive EDX-220-21-E
Fig. 3. Exhaust evaporator.
Table 4
Measurement devices accuracy.
Input Type Range Accuracy Unit
Engine speed Schenck D900-1e 0e6500 ±2 rpm
Engine torque Schenck D900-1e 4000e4000 ±8Nm
Expander speed HBM T40B 0e20,000 ±10 rpm
Expander torque HBM T40B 500e500 ±0.25 Nm
Fuel ow AVL 730 0e150 ±0.9 kg/h
Mass ow Micro Motion F025S 0e100 ±0.2 g/s
Cycle high pressure WIKA A-10 0e60 ±0.6 bar(g)
Cycle low pressure WIKA A-10 0e6±0.06 bar(g)
Exhaust pressure WIKA A-10 0e2.5 ±0.03 bar(g)
Pressure drop Yokogawa EJA110E 0e5000 ±2.75 Pa
Temperature RS Pro Type K 75e1100 ±1.5
C
J. Rijpkema, O. Erlandsson, S.B. Andersson et al. Energy 238 (2022) 121698
4
approximated in Eq. (4) using technical data from the manufacturer
[55].
_
V
corr
¼0:0001N
pmp
þ1:2 (4)
The pump power can be calculated with an estimated efciency:
_
W
pmp
¼_
m
pmp
ðh
pmp;out
h
pmp;in
Þ.
h
pmp
(5)
To validate the pump model, measurements were conducted at
a low (set 1) and a high (set 2) pressure. In both cases, no heat was
added to the system, the pump bypass valve was fully closed, the
expander was not running, and the expander bypass valve was used
to control the pressure. The results of the experiments and simu-
lations are shown in Fig. 6, indicating that a good agreement was
achieved.
3.3. Pump bypass valve
The pump bypass valve is modeled as an incompressible ow
valve, as shown in Eq. (7). The discharge coefcient (C
d
) and valve
area (A) can be combined into an effective area (A
bpv
).
_
m
bpv
¼C
d
Affiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2
r
in
ðp
in
p
out
Þ
p¼A
bpv
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2
r
in
ðp
in
p
out
Þ
p(6)
Although the pump bypass mass ow ( _
m
bpv
) is not known, the
evaporator mass ow ( _
m
evap
) was measured and can be subtracted
from the pump mass ow ( _
m
pmp
), available from the pump model:
_
m
bpv
¼_
m
pmp
_
m
evap
(7)
The effective area is a function of the valve position (x
bpv
) and is
calibrated using the data from experimental set 3. The calibrated
model consists of a combination of two linear functions of which
the coefcients are shown in Table 8.
The resulting effective area as a function of the valve position is
shown on the left on Fig. 7, together with the corresponding
experimental data. During these experiments no heat was added to
the system, the pump bypass valve position was controlled, the
expander was not running, and the expander bypass valve was held
at a xed position. The corresponding results for the mass ow is
shown on the right of Fig. 7, along with results obtained under
Table 5
Numbering, quantity of experimental points (Qty.) and description of the experimental sets.
Set Qty. Description
1. 7 Cold system, no expander, low pressure, pump validation
2. 6 Cold system, no expander, high pressure, pump validation
3. 13 Cold system, no expander, pump bypass valve calibration
4. 16 Cold system, no expander, pump bypass valve validation
5. 28 Hot system, no expander, engine results, evaporator calibration, condenser results
6. 41 Hot system, expander calibration, cycle validation
Table 6
Measurement error for the derived quantities.
Quantity Symbol Max. Error
Engine mass ow _
m
eng
±5.9%
Pump power _
W
pmp
±9.4%
Bypass valve effective area A
bpv
±4.5%
Evaporator heat transfer rate _
Q
evap
±6.8%
Expander lling factor 4
f, is
±6.0%
Expander efciency
h
exp
±8.9%
Expander power _
W
exp
±8.2%
Condenser heat transfer rate _
Q
cond
±7.9%
Fig. 4. Engine speed-torque map showing the measured engine exhaust mass ow.
The measurements were corrected by a factor 0.75 based on an estimated error of the
measured values.
Fig. 5. Engine speed-torque map showing the measured engine exhaust outlet
temperature.
Table 7
Pump specications [55].
Maximum pressure p
max
100 bar
Displacement volume V
pmp
0.002 L
J. Rijpkema, O. Erlandsson, S.B. Andersson et al. Energy 238 (2022) 121698
5
different conditions (set 4) that were used for model validation.
Good agreement with the experiments is observed for both for the
calibration and validation data sets.
3.4. Evaporator
The exhaust evaporator is a cross-counter ow plate heat
exchanger with ns; its heat transfer area is depicted schematically
in Fig. 8. The exhaust ow side has one pass and 36 channels, while
the water side has three passes and 35 channels. The dimensions of
the heat exchanger are listed in Tables 9 and 10. The model only
accounts for heat transfer; the pressure drop over the evaporator is
ignored.
Heat transfer is modeled using n-specic equations. The geo-
metric parameters of the ns are computed using the values pre-
sented in Table 10 and the following equations [56]:
D
h
¼4s
f
bL
f
2ðs
f
L
f
þbL
f
þt
f
bÞþt
f
s
f
(8)
A
flow
¼s
f
h
f
(9)
A
base
¼2ðs
f
L
f
þt
f
ðs
f
t
f
Þ.2Þ(10)
A
fin
¼2ðh
f
L
f
þh
f
t
f
Þ(11)
For each pass and channel the number of ns for the working
uid and exhaust can be calculated:
n
f;x;wf
¼L
p
f
;n
f;y;wf
¼H
L
f
(12)
n
f;x;exh
¼L
L
f
;n
f;y;exh
¼H
p
f
(13)
As a result, the total cross-sectional area (or total ow area) for
the ow can be calculated:
Fig. 6. The mass ow as a function of the pump speed and pump outlet pressure.
Table 8
Pump bypass valve coefcients.
x
bpv
% 0 1 2 7 100
A
bpv
mm
2
0 0.25 0.35 1 2.65
Fig. 7. Pump bypass valve effective area (left) and mass ow (right) as a function of the bypass valve position.
Fig. 8. Schematic depiction of the heat transfer area of the exhaust evaporator.
Table 9
Heat exchanger dimensions.
Heat exchanger length L144 mm
Heat exchanger width W241 mm
Heat exchanger height H247 mm
Channel height b3.00 mm
Plate thickness t0.40 mm
Fluid-specic wf exh
Number of channels n
ch
35 36 e
Number of passes n
pass
31 e
Number of ns in x-dir. n
f, x
96.0 45.4 e
Number of ns in y-dir. n
f, y
77.8 164.7 e
Number of ns n
f
7468 7468 e
Total ow area A
c
4309 22,808 mm
2
J. Rijpkema, O. Erlandsson, S.B. Andersson et al. Energy 238 (2022) 121698
6
A
c;wf
¼n
ch;wf
n
f;x;wf
n
pass;wf
A
flow
(14)
A
c;exh
¼n
ch;exh
n
f;y;exh
n
pass;exh
A
flow
(15)
The mass ux (G) for the exhaust and water sides is calculated
by dividing the mass ow by the corresponding cross-sectional
area (G¼_
m=A
c
). With these denitions and the associated ther-
modynamic and transport properties, the Reynolds number (Re),
the Prandtl number (Pr), and the heat transfer coefcient (
a
) can be
determined using the Nusselt (Nu) number. For single-phase heat
transfer, the following expression for the Nusselt number [56]is
used:
Nu ¼jRePr
1=3
(16)
Where:
j¼0:652Re
0:540
s
f
h
f
!
0:154
t
f
L
f
!
0:150
t
f
s
f
!
0:068
,2
41þ5:269,10
5
Re
1:34
s
f
h
f
!
0:504
t
f
L
f
!
0:456
t
f
s
f
!
1:06
3
5
0:1
(17)
The two-phase heat transfer coefcient is the sum of the
nucleate boiling (
a
nb
) and convective (
a
cv
) components [56]:
a
¼
a
nb
þ
a
cv
(18)
The nucleate boiling component is a function of the heat ux ( _
q)
for each element, the molecular weight (M
w
), and the reduced
pressure (p
crit
), and is dened as:
a
nb
¼55_
q
2=3
M
1=2
w
p
p
crit
0:225
log
10
p
p
crit

0:55
(19)
The convective component is obtained from the saturated liquid
heat transfer coefcient (
a
l
), which is computed using Eq. (16) with
saturated liquid properties. This coefcient is then multiplied by a
factor (F) that depends on the steam quality (x) and the saturated
liquid and vapor densities (
r
l
,
r
v
) and viscosities (
m
l
,
m
v
):
a
cv
¼F
a
l
(20)
F¼1þ28
X
tt
0:372
(21)
X
tt
¼1x
x
0:9
r
v
r
l
0:5
m
v
m
l
0:1
(22)
To solve the heat transfer equations, the heat exchanger is dis-
cretized into a set of elements as shown in Fig. 9. A low resolution
model is compared to a high resolution TitanX model whose set-
tings are shown in Table 11. The heat transfer surface (A
s
) is the
combined base (A
base
) and n(A
n
) surface:
A
s;el
¼A
base;el
þA
fin;el
(23)
The heat transfer is calculated for all channels based on the total
surface area for each element:
A
s;el;tot
¼n
ch;wf
A
s;el
(24)
For each element the heat transfer can be calculated:
_
Q
el
¼U
el
A
s;el;tot
ðT
exh;el
T
wf;el
Þ(25)
The overall heat transfer coefcient (U) consists of the sum of
the separate contributions:
1
U¼1
a
wf
þt
w
l
w
þ1
a
exh
z1
a
wf
þ1
a
exh
(26)
The experimental and simulation results of the evaporator heat
transfer rate and outlet temperature are shown in Fig. 10, based on
experimental set 5 from Table 5. Experiments were performed at
every engine operating point shown in Figs. 4 and 5 (A25-C100). In
these experiments, the pump and pump bypass valve position were
controlled to maintain a constant mass ow at each operating
point. The expander was not running and the expander bypass
valve was used to control the pressure. Both the low resolution
model and the TitanX model exhibit good agreement with the
experiments. However, it should be noted that in most cases the
available heat from the exhaust gases was so large that the water
was superheated to a temperature close to the exhaust gas inlet
temperature. Increasing the mass ow would reduce the evapo-
rator outlet temperature of the water, but the magnitude was
difcult to control experimentally. Because of the relatively small
mass ows and high latent heat of water, small deviations in mass
ow caused large deviations in the evaporator outlet temperature.
Another possible source of error is that the temperatures were
measured at a single location, slightly downstream of the evapo-
rator outlet. This could cause variations between the experimental
values due to heat loss and local effects, leading to deviations from
the model values. However, since the heat transfer to the working
uid is the most important for the prediction of the cycle perfor-
mance, the model can still be used in the cycle simulations.
Table 10
Fin geometry.
Spacing s
f
1.35 mm
Thickness t
f
0.15 mm
Strip ow length L
f
3.175 mm
Pitch p
f
1.50 mm
Effective channel height h
f
2.85 mm
Hydraulic diameter D
h
1.791 mm
Flow area A
ow
3.848 mm
2
Base area A
base
8.753 mm
2
Fin area A
n
18.95 mm
2
Fig. 9. Discretization of the exhaust evaporator.
J. Rijpkema, O. Erlandsson, S.B. Andersson et al. Energy 238 (2022) 121698
7
3.5. Expander
The expander is an uniow reciprocating piston expander with
two cylinders of which the relevant geometrical specications are
shown in Table 12.
The expander model is based on a semi-empirical model for
volumetric expanders [57] which is schematically shown in Fig. 11.
The semi-empirical model consists of thermodynamic equations
with tuning parameters that are determined by calibrating them
against experimental results. Using these parameters, the de-
viations from ideal expander performance caused by pressure
drops, leakage, heat losses, and mechanical losses can be deter-
mined. More details on the modeling, results, and validation of the
expander were presented in a previous publication [58].
The operation of the expander is characterized by three main
performance factors: the isentropic lling factor (4
f, is
), the
expander efciency (
h
exp
), and the isentropic effectiveness (ε
is
). All
three factors depend on the process conditions and expander
operation. The isentropic lling factor (4
f, is
) is used to predict how
much the expander mass ow will deviate from ideal conditions
under isentropic outlet conditions. For a piston expander, the
theoretical ow is equal to product of the expander speed (N
exp
)
and the difference between the available mass at inlet valve closing
(
r
exp, in
f
a
V
exp
) and the trapped mass at exhaust valve closing (
r
exp,
out
f
p
V
exp
). In all tested conditions, the expansion ended in the two-
phase region. Therefore, the outlet conditions are based on the
isentropic conditions (
r
exp, out, is
f
p
V
exp
). This leads to the following
expression:
_
m
exp
¼
r
exp;in
f
a
r
exp;out;is
f
p
4
f;is
V
exp
N
exp
60 (27)
To predict the shaft power output of the expander ( _
W
exp
), the
expander efciency is used:
_
W
exp
¼
h
exp
_
m
exp
ðh
exp;in
h
exp;out;is
Þ(28)
For the expander considered in this paper, not only the shaft
power output, but also the leakage and heat loss effects were sig-
nicant. Therefore, the isentropic effectiveness (ε
is
) is introduced,
which is used to calculate the expander outlet enthalpy:
h
exp;out
¼h
exp;in
ε
is
ðh
exp;in
h
exp;out;is
Þ(29)
The resulting model outputs and the corresponding experi-
mental values are shown on the left of Figs. 12 and 13, taken from
experimental set 6. Experiments were performed at four engine
operating points (A25, HW, A50, and B25). In the experiments, the
pump and pump bypass valve position were controlled to provide a
constant mass ow at each engine operating point, the expander
speed was varied, and the expander bypass valve was closed. The
results show that the expander mass ow is well captured by the
isentropic lling factor. However, the shaft power is overpredicted
for low pressure ratios (and corresponding low power outputs) and
underpredicted at higher pressure ratios (and corresponding high
power outputs). Since the expander efciency is not only depen-
dent on the pressure ratio, variations between experimental values
for similar pressure ratios occur. Other important physical quanti-
ties include the expander speed, the cycle mass ow, and the
expander inlet temperature. Deviations between model and
experimental values are mainly attributed to the expansion in the
two-phase region, high leakage rate in the expander, and the
change in lubrication properties over time. These topics are dis-
cussed in more detail in a separate publication [58].
3.6. Condenser
The condenser is modeled as a heat sink only because neither
detailed information on its geometry nor ow measurements on
Table 11
Heat exchanger geometry for each element.
Model TitanX
Number of els. in x-dir. n
el, x
69e
Number of els. in y-dir. n
el, y
310e
Number of ns n
f, el
415 83 e
Base area A
base, el
3631 726 mm
2
Fin area A
n, el
7864 1573 mm
2
Heat transfer area A
s, el
11,495 2299 mm
2
Fig. 10. Evaporator heat transfer rate (left) and outlet temperature (right) as a function of the mass ow.
Table 12
Expander specications.
Supply cut-off f
a
0.16 e
Exhaust cut-off f
p
0.78 e
Displaced volume V
exp
0.8 L
Fig. 11. Schematic depiction of the semi-empirical model used for the expander [58].
J. Rijpkema, O. Erlandsson, S.B. Andersson et al. Energy 238 (2022) 121698
8
the process water used for cooling were available. Assuming no
leakage of the working uid, the modeled condenser ow was
considered to be equal to the evaporator ow. This gives the
following relations for the condenser:
_
m
cond
¼_
m
evap
(30)
_
Q
cond
¼_
m
cond
ðh
cond;out
h
cond;in
Þ(31)
Fig. 14 shows the experimental results of the condenser heat
transfer rate from experimental set 5 in Table 5 together with the
corresponding outlet temperatures. Experiments were performed
at all of the engine operating points shown in Figs. 4 and 5 (A25-
C100). In these experiments, the pump and pump bypass valve
position were controlled to provide a constant mass ow at each
engine operating point. The expander was not running and the
expander bypass valve was used to control the pressure.
3.7. Rankine cycle
Simulations using the full cycle model incorporating all of the
calibrated component models discussed in the preceding sections
were performed in MATLAB [59] using uid maps generated from
the CoolProP [60] database. Fig. 15 shows the comparison of these
simulation results to data from experiment set 6 in Table 5. Ex-
periments were performed at four engine operating points (A25,
HW, A50, and B25). The pump speed and pump bypass valve po-
sition were controlled to provide a constant mass ow, the
expander speed was varied, and the expander bypass valve was
closed.
The comparison of the model output to the experimental values
shows that the pump outlet pressure is well captured. Deviations in
the mass ow are mostly due to the sensitivity of the pump bypass
valve model; small pressure changes cause large changes in the
predicted ow through the bypass valve, leading to poor agreement
between the model and experiment. Because the evaporator heat
Fig. 12. The isentropic lling factor (left) and mass ow (right) of the expander as functions of the expander speed. Symbols indicate experimental results and lines indicate model
outputs.
Fig. 13. The efciency (left) and shaft power (right) of the expander as functions of the expander speed. Symbols indicate experimental results and lines indicate model outputs.
Fig. 14. Condenser heat transfer rate (left) and outlet temperature (right) as functions of the mass ow.
J. Rijpkema, O. Erlandsson, S.B. Andersson et al. Energy 238 (2022) 121698
9
transfer rate is proportional to the mass ow, the same effect is
visible there. Also visible is the overprediction of the evaporator
outlet temperature. Since no heat loss was considered in the
models, the evaporator outlet temperatures of the working uid
almost reached the exhaust gas temperature. In reality, heat losses
would lead to slightly lower temperature, explaining the deviations
between simulations and experiments. Finally, the expander
output power is signicantly overpredicted at low power outputs
and underpredicted at higher power outputs; the reasons for this
are discussed in more detail in a separate publication [58].
The main source of error between the experimental and simu-
lated values of the mass ow is the bypass valve model. A smaller
pump in the experimental setup would possibly eliminate the need
for a bypass valve. Alternatively, a more sophisticated model or
calibration method for the bypass valve could improve the t be-
tween experiments and simulations. Using a different working uid
could also help improve the t. A uid with a smaller latent heat
would mean higher mass ows to extract the available heat.
Simultaneously, the calibration and validation for the heat
exchanger and the expander models could be improved. Another
improvement would be the addition of pressure drop and heat loss
correlations to the models. Although the gures show that the
deviations between simulation and experimental results can be
signicant, the simulation results are based on physical models and
the general trend is well-captured. This means that these models
are a valid tool to compare the performance of different working
uids for operational range of the engine, which is done in the
subsequent sections.
4. Simulation setup
The validated cycle model allows for predicting the performance
of the WHR system under conditions outside the experimentally
tested range (e.g. over a driving cycle) and with different working
uids. For this purpose, the bypass valve is removed from the
original model, giving the cycle schematically depicted in Fig. 13.To
obtain the desired mass ow, the pump is allowed to operate at
speeds outside the range specied by the manufacturer.
4.1. Working uids
To evaluate the performance of the WHR system, simulations
were performed with two additional working uids: cyclopentane
and ethanol. These uids were selected based on their promising
thermodynamic performance in heavy-duty engine applications
[17,61]. Previous studies by the authors [6,16,62], where different
heat sources from the engine were evaluated for many different
working uids, also showed good thermodynamic performance for
these two uids. Additionally, they are environmentally friendly,
relatively non-toxic, and non-corrosive, although ammability is a
concern for both of them. Table 13 lists a number of important
properties for the three working uids.
4.2. Driving cycle
The model was calibrated against experimental data obtained in
an engine test cell. However, temperatures under driving condi-
tions are usually lower due thermal inertia and heat loss in the
aftertreatment systems and exhaust piping. Therefore, the input
conditions for the simulations were taken from a representative
driving cycle for a 40 tonne EU6 Scania long haul truck driving on a
European road. The vehicle speeds and road gradients for this
driving cycle are shown in Fig. 16.
The exhaust outlet conditions (temperature and mass ow)
during the driving cycle are divided into a four-by-four grid, as
Fig. 15. Comparison of full cycle model outputs and experimental results.
Table 13
Properties of the selected working uid.
Fluid MW p
crit
T
crit
GWP ODP Type
- kg/kmol bar
C-
Cyclopentane 70.1 45.7 239 0 0 isentropic
Ethanol 46.1 62.7 240 0 0 wet
Water 18.0 220 374 0 0 wet
J. Rijpkema, O. Erlandsson, S.B. Andersson et al. Energy 238 (2022) 121698
10
shown in Fig. 18. The percentages in the grid represent the relative
duration of these conditions during the driving cycle. The green
values indicate the duration for a positive torque on the engine and
the red values indicate negative torque. The center points for all 16
elements were used as inputs for the steady-state simulations.
Transient effects during the driving cycle were not taken into
account.
4.3. Cycle performance
To run the simulations, a number of inputs and constraints were
specied, as summarized in Table 14. The exhaust mass ow,
temperature, and pressure were taken from the 16 previously
dened operating conditions. Depending on whether the engine
torque was positive or negative, the expander was coupled directly
to the engine via a mechanical coupling or to an electrical gener-
ator. A subcooling temperature difference of 5 K was set to prevent
vapor entering the pump. The pump mechanical efciency, elec-
trical generator efciency, and the efciency of the mechanical
coupling between expander and engine were taken to be 0.50, 0.85,
and 0.98, respectively. Only subcritical conditions were taken into
account. The evaporator outlet temperature was limited to avoid
thermal instability of the working uid and overheating of sus-
pended oil. To allow for temperature control, a minimum and
maximum superheating temperature difference were set. The
range of expander speeds was based on the specications from the
manufacturer [63]. The pump speed was not limited by these
specications; instead it was set to give the highest possible mass
ow. No pressure drops in the components were considered and
component heat losses other than the expander heat loss were
ignored. A golden section search was performed to nd the
expander speed providing the maximum power output for the
stated inputs and constraints.
Cycle performance is evaluated based on the net (shaft) power
and thermodynamic efciency:
_
W
net
¼
_
W
exp
_
W
pmp
(32)
h
th
¼
_
W
net
_
Q
evap
(33)
The performance in the driving cycle is estimated using the
denitions expressed in Eqs. (34) and (35). At positive engine tor-
que (
t
eng
), the expander power is directly provided to the engine,
while at negative engine torque the expander power is converted
into electrical power.
_
W
pmp;el
¼
_
W
pmp
h
el
(34)
_
W
exp
¼(_
W
exp;mech
¼
h
mech
_
W
exp
;if
t
eng
>0
_
W
exp;el
¼
h
el
_
W
exp
;if
t
eng
0(35)
5. Results and discussion
5.1. Steady-state performance
Steady-state simulations using the cycle model with water as
the working uid were performed for 16 engine operating points
with exhaust mass ows ( _
m
exh
) ranging from 150 to 450 g/s and
exhaust outlet temperatures (T
exh, out
) between 260 and 320
C, as
previously presented in Fig. 18 and Table 14. In the following dis-
cussion, these will be designated with an M for mass ow and a T
for temperature. Thus, M150T300 corresponds to an exhaust mass
ow of 150 g/s and an outlet temperature of 300
C. The pump
Fig. 16. Schematic of full cycle model.
Fig. 17. Representative driving cycle for a long haul truck on an European road.
Fig. 18. Relative time distributions of the exhaust mass ow and temperature over the
driving cycle. Green values indicate positive torque and red negative torque. The total
drive cycle duration (
D
t
dc
) is 2748 s. (For interpretation of the references to colour in
this gure legend, the reader is referred to the Web version of this article.)
J. Rijpkema, O. Erlandsson, S.B. Andersson et al. Energy 238 (2022) 121698
11
outlet pressures and mass ows at four engine operating points for
a range of expander speeds are shown in Fig. 19. In these cases, the
evaporator superheating temperature difference was kept within
the constraints specied in Table 14. At each engine operating
point, the exhaust temperature was kept constant while varying
the mass ow, effectively changing the heat available for recovery.
As expected, the pump outlet pressure decreased as the
expander speed increased and increased as the available exhaust
heat increased (represented by the engine operating points). An
increase in the available exhaust heat allows for higher cycle mass
ows, as shown on the right of Fig. 19. The same effect was
observed when increasing the expander speed. Increasing the
expander speed (and thus reducing the evaporating pressure of
water in the cycle, as explained further on; see Fig. 22) made it
possible to recover more heat from the exhaust gases. The resulting
net power, shown on the left of Fig. 20, depends on the required
pump power, the amount of heat transferred from the heat source,
and the expander power obtained. The pump power is determined
by the pump efciency, the pressure difference over the pump,and
the mass ow. The amount of heat transferred from the source to
the cycle is a function of the exhaust mass ow and temperature, as
well as the cycle temperature, pressure, and mass ow. The
expander power depends on how effectively the recovered heat is
transformed into power, which is shown on the right of Fig. 20.
Because of the interaction between the pump power, recovered
heat, and expander power, there is no single optimal expander
speed that maximizes the power output for all engine operating
points. For the lower exhaust mass ows, the maximum net power
is around 1.1 kWand is achieved at a relatively low expander speed
of around 900 rpm. As the exhaust mass ow increases, both the
power output and the optimal expander speed increase, with
maxima of 4.2 kW and 2800 rpm, respectively.
Simulations were performed to obtain the maximum power
output for the 16 engine operating points with water as the
working uid. The expander speed was varied at each operating
point to obtain the maximum net power output, which is shown on
the left of Fig. 21. The values shown at the edges of this gure were
obtained by linear extrapolation. Depending on the exhaust mass
ow and temperature, the recoverable net power ranges from 0 to
8 kW. Another important aspect for automotive applications is the
amount of heat that must be rejected to allow condensation of the
working uid, which is shown on the right of Fig. 21. When using
exhaust gases as a heat source, this heat must be either transferred
to the coolant and rejected in the coolant radiator or rejected
directly via a separate radiator. The results of the simulations show
that the heat transfer rate in the condenser can be as high as 60 kW.
Simulations using the same cycle components were also per-
formed with cyclopentane and ethanol as the working uid, and
the results all selected working uids are shown in Table 15. The
condensation pressure was set at 1.1 bar for all uids, resulting in a
different condensation temperature for each uid. A lower
condensation temperature means a smaller temperature difference
between the working uid and the ambient temperature, making it
more difcult to reject excess heat. In a practical system, this could
lead to increased power consumption by the cooling fan, which
would reduce the net power output of the system. This effect is not
taken into account here.
The results from Table 15 show that the highest power output
was obtained with cyclopentane, then ethanol, and nally water.
These differences in the performance can be explained by consid-
ering the M150T300 engine operating point, for which the results
are shown in Table 16.
The power outputs for cyclopentane and ethanol were higher
than for water because of higher mass ows and expander
Table 14
Cycle inputs and constraints.
Inputs
Exhaust gas mass ow _
m
exh
150e450 g/s
Exhaust gas inlet temperature T
exh, in
260e320
C
Exhaust gas inlet pressure p
exh, in
1.03e1.06 bar
Pump inlet subcooling temperature
D
T
sub
5K
Pump mechanical efciency
h
pmp
0.50 e
Electrical generator efciency
h
el
0.85 e
Mechanical coupling efciency
h
mech
0.98 e
Constraints
Pump outlet pressure p
pmp, out
10 ep
crit
bar
Evaporator outlet temperature T
evap, out
n/a e260
C
Evaporator outlet superheating temp.
D
T
sup
10e30 K
Expander speed N
exp
500e3500 rpm
Pump speed N
pmp
150e6000 rpm
Fig. 19. Pump outlet pressure (left) and mass ow rate (right) for water as functions of the expander speed at different engine operating points.
J. Rijpkema, O. Erlandsson, S.B. Andersson et al. Energy 238 (2022) 121698
12
efciencies. The higher mass ows of cyclopentane and ethanol are
due to the lower latent heat of these uids and to a better thermal
match between the heat source and the cycle. This is visualized in
the heat transfer-temperature (QT) diagrams for the different
working uids at the M150T300 engine operating point, which are
presented in Fig. 22. For cyclopentane, the temperature slope dur-
ing preheating matches the temperature prole of the heat source,
meaning that the heat transfer is not limited by the evaporating
temperature of the working uid and the maximum heat can be
extracted for this specic heat exchanger geometry. For ethanol and
especially for water, less heat can be extracted and a lower power
output is achieved.
The corresponding temperature-entropy (Ts) diagrams are
shown in Fig. 23. This shows that the expansion ends in the two-
phase region in the cases of ethanol and water whereas for cyclo-
pentane it ends in the superheated region. For most expanders,
Fig. 20. Net shaft power (left) and expander efciency (right) for water as functions of the expander speed at different engine operating points.
Fig. 21. Net shaft power (left) and condenser heat transfer rate (right) for water as a function of the exhaust mass ow and temperature. The values at the edges were linearly
extrapolated.
Fig. 22. QT-diagrams of cyclopentane, ethanol, and water for the M150T300 engine operating point.
Table 15
Range of cycle conditions for the 16 engine operating points.
Fluid _
m
pmp
p
evap
T
exp, in
N
exp
p
cond
T
cond
_
Q
cond
_
W
net
h
th
- g/s bar
C rpm bar
CkW kW %
Cyclopent. 38.6e142 19.4e30.9 194e225 1285e3500 1.1 52 17.4e75.0 1.8e9.6 7.7e11
Ethanol 18.6e74.4 20.9e30.1 193e232 861e2774 1.1 80 14.6e66.3 1.0e7.8 5.9e9.8
Water 5.16e22.6 15.1e26.5 210e255 700e3139 1.1 102 10.2e45.8 0.5e5.7 3.7e10
J. Rijpkema, O. Erlandsson, S.B. Andersson et al. Energy 238 (2022) 121698
13
operation in the superheated region is preferable because it pre-
vents droplets from damaging the expander. Additionally, when
using cyclopentane, a recuperator could be added to improve the
power output or reduce the condenser load. Although the Ts-
diagrams suggest that the expander is operating at isentropic ef-
ciency, this is not the case. During the expansion, not all of the
available energy is converted to power; a signicant portion is lost
as heat, which is why two denitions for the expander power and
energy change are necessary, as dened in Eq. (28) and Eq. (29).
5.2. Driving cycle performance
The exhaust ow and temperature over the driving cycle pre-
sented in Fig. 17 were used as inputs for the steady-state simula-
tions of the three selected working uids. The resulting net power
outputs over the driving cycle of these uids are shown in Fig. 24.
In accordance with the steady-state simulations, the best per-
formance was obtained with cyclopentane, followed by ethanol and
then water. To estimate performance over the whole driving cycle,
the results were numerically integrated using a timestep (
D
t)of1s:
W¼X
n
i¼1
_
W
i
D
t(36)
The integrated results for Eqs. (31), (34) and (35) are shown in
Table 17, both in absolute and relative terms. The relative perfor-
mance is obtained by dividing the absolute result by the total work
done by the engine during the driving cycle (W
dc
¼333 MJ). The
results show that the WHR system can recover a signicant amount
of energy, corresponding to as much as 3.37% of the total engine
energy requirement. This relative recovery can be roughly trans-
lated into fuel consumption reduction, assuming that the increased
backpressure due to the exhaust evaporator does not affect the
engine efciency. The results also show that the recovered elec-
trical work is comparable to the electrical work done by the pump,
although it must be noted that the efciencies of storing and
extracting power from the battery are not included. Additionally,
the table shows that the pump work is much higher for cyclo-
pentane than for the other uids because of the higher mass ow in
the cycle. However, this is more than offset by the increase in
expander power.
Even though a driving cycle was used to evaluate the perfor-
mance of the working uids over the operational range of the en-
gine during actual driving conditions, transient effects were
ignored in this paper. The cycle components are assumed to react
instantaneously to changes in the exhaust ow and temperature. In
reality, the performance of the components will be affected by
inertia during transient operation, with the thermal inertia in the
heat exchangers being dominant [64]. This has important impli-
cations for the control of the system, as superheated conditions at
the inlet of the expander should be ensured [14]. Another study
[65] showed in a comparison between a steady state and transient
model that a transient model will lead to lower fuel savings,
although the resulting ranking remains the same. The expander
coupling is another point of discussion. If the expander is me-
chanically coupled to the engine, its speed will be determined by
Table 16
Cycle conditions for the M150T300 engine operating point.
Fluid _
m
pmp
p
evap
T
exp, in
N
exp
p
cond
T
cond
_
Q
cond
_
W
net
h
th
- g/s bar
C rpm bar
CkWkW%
Cyclopentane 52.9 21.6 199 1546 1.1 52 27.7 2.75 8.7
Ethanol 25.2 23.4 214 1023 1.1 80 20.9 1.81 6.9
Water 7.45 18.7 228 1023 1.1 102 14.7 1.09 6.0
Fig. 23. Ts-diagrams of cyclopentane, ethanol, and water at the M150T300 engine operating point.
Fig. 24. Net shaft power for multiple uids during the driving cycle.
Table 17
Driving cycle performance with a total engine work requirement of 333 MJ.
Fluid Q
cond
W
pmp, el
W
exp, mech
W
exp, el
W
*
net
MJ MJ (%) MJ (%) MJ (%) MJ (%)
Cyclopentane 100 1.62 (0.49) 11.7 (3.50) 1.17 (0.35) 11.2 (3.37)
Ethanol 86.1 0.90 (0.27) 8.42 (2.53) 0.69 (0.21) 8.21 (2.46)
Water 57.1 0.20 (0.06) 5.02 (1.51) 0.37 (0.11) 5.19 (1.56)
*W
net
¼W
exp, mech
þW
exp, el
W
pmp, el
.
J. Rijpkema, O. Erlandsson, S.B. Andersson et al. Energy 238 (2022) 121698
14
the engine speed and a predetermined gear ratio. However, the
driving cycle results presented here were obtained under the
assumption that the expander could operate at its optimal speed for
which the corresponding power output were determined in the
earlier steady-state simulations. The results in this section allow for
comparison of different working uids on the basis of their ther-
modynamic performance. In practice, however, the selection of the
best working uid is also subject to other constraints such as costs,
component sizing, and environmental impact.
6. Conclusions
Experimental investigations were performed to evaluate the
performance of a Rankine system with water for WHR from the
exhaust gases of a heavy-duty engine. The results of these experi-
ments constitute one of the main contributions of this paper.
Additionally, models of the relevant cycle components were
developed and then calibrated and validated against the experi-
mental data. These models provided more detailed overview of the
physical processes occurring within each component. This allowed
for predictions of the performance of these components under
conditions outside the experimental range and when using
different working uids. The component models were combined to
create model of the full cycle, allowing the performance of the
Rankine system to be simulated over a typical long haul truck
driving cycle. The main results and conclusions obtained during
this work were:
CExperimental results were obtained for a wide range of en-
gine operating conditions. The experiments were divided
into six distinct experimental sets and their results were
used to calibrate and validate models of the main compo-
nents of the Rankine cycle, i.e. the pump, pump bypass valve,
evaporator, expander, and condenser. Experimental mea-
surements of the expander shaft power were performed at
four different engine operating points (A25, HW, A50, B25).
The expander power ranged from 0.2 to 3 kW, corresponding
to 0.2e2.5% of the engine power.
CSteady-state simulations of the Rankine cycle with water as
the working uid exhibited good agreement with the
experimentally determined mass ow and evaporator heat
transfer, but the expander power was overpredicted at low
expander powers and underpredicted at high powers. Sim-
ulations performed at 16 engine operating points gave net
power outputs between 0.5 and 5.7 kW, and the optimal
expander speed was found to be dependent on the engine
operating point. The added heat needing to be rejected in the
condenser was between 10 and 46 kW. These values can be
extrapolated to obtain results for the full range of operating
conditions, yielding net power outputs between 0 and 8 kW
and condenser heat transfer rates between 0 and 60 kW.
CTo evaluate the performance of different working uids in
the studied WHR system, simulations were also performed
with cyclopentane and ethanol as the working uids. The
results indicated that the evaporating pressures and
expander inlet temperatures for these uids were similar to
those of water, but that they had higher mass ows in the
cycle. The increased mass ows were a result of the lower
latent heat and a better thermal match with the heat source,
allowing for more heat transfer between source and cycle.
Because of the higher ows and expander efciencies,
cyclopentane and ethanol outperformed water, providing
net power outputs between 1.8 and 9.6 kW and 1.0 and
7.8 kW, respectively.
CThe steady-state results for the three working uids were
used to simulate the performance over a typical driving cycle
of a long haul truck. Although transient effects were not
taken into account and the expander speed was not
controlled by the engine speed (as would be the case in a real
system due to the mechanical coupling), the results still
allow for a comparison between the thermodynamic per-
formance of the systems with the different working uids.
The total recovered energy during the driving cycle was 11.2,
8.2, and 5.2 MJ for cyclopentane, ethanol, and water,
respectively, corresponding to recoveries of 3.4, 2.5, and 1.6%
relative to the total energy requirement of the engine.
Declaration of competing interest
The authors declare that they have no known competing
nancial interests or personal relationships that could have
appeared to inuence the work reported in this paper.
Acknowledgments
This research was made possible by funding provided by the
Strategic Vehicle Research and Innovation Programme (FFI) of the
Swedish Energy Agency. The authors would like to thank the
partners in the WHR project: Gnutti Carlo, IAV, Lund University,
Scania, TitanX, Volvo Cars, and Volvo Group.
Nomenclature
Aarea (m
2
)
C
d
discharge coefcient ()
hspecic enthalpy (J/kg)
MW molecular weight (kg/kmol)
_
mmass ow rate (kg/s)
Nrotational speed (rpm)
ppressure (Pa)
_
Qheat transfer rate (W)
sentropy (J/kg/K)
ttime (s)
Ttemperature (K)
Vvolume (m
3
)
_
Vvolume ow rate (m
3
/s)
_
Wpower (W)
Greek symbols
h
efciency ()
εeffectiveness ()
4
f
lling factor ()
r
density (kg/m
3
)
t
torque (Nm)
Subscripts
amb ambient
bpv bypass valve
cond condenser
cool coolant
corr correction
crit critical
Subscripts (continued)
el electrical
eng engine
evap evaporator
exh exhaust
J. Rijpkema, O. Erlandsson, S.B. Andersson et al. Energy 238 (2022) 121698
15
exp expander
is isentropic
mech mechanical
pmp pump
sh shaft
sub subcooled
sup superheated
th theoretical/thermodynamic
Abbreviations
CAC
charge air cooler
BPV
bypass valve
EGR
exhaust gas recirculation
EGRC
exhaust gas recirculation cooler
ESC
European stationary cycle
HD
heavy-duty
HDD
heavy-duty Diesel
HW
highway
GHG
greenhouse gas
GWP
global warming potential
ODP
ozone depletion potential
ORC
organic Rankine cycle
SV
safety valve
WHR
waste heat recovery
References
[1] International Energy Agency. Energy technology Perspectives 2020. 2020.
[2] European Environment Agency (EEA). National emissions reported to the
UNFCCC and to the EU greenhouse gas monitoring mechanism. 2017. https://
www.eea.europa.eu/ds_resolveuid/a6e1bc85fbed4989b0fd6739c443739a.
[Accessed 15 December 2020].
[3] European Parliament, Council of the European Union. Regulation (EU) 2019/
1242 - Setting CO2 emission performance standards for new heavy-duty ve-
hicles. 2019.
[4] Hofer F, Gruber W, Raser B, Theißl H. Technology scenarios for fullling future
EU CO2 targets for commercial vehicle eets. ATZheavy duty worldwide
2020;13(1).
[5] Pischinger S, Schaub J, Aubeck F, van der Put D. Powertrain concepts for
heavy-duty applications to meet 2030 CO2 regulations. ATZheavy duty
worldwide 2020;13(3).
[6] Rijpkema J, Munch K, Andersson SB. Thermodynamic potential of twelve
working uids in Rankine and ash cycles for waste heat recovery in heavy
duty diesel engines. Energy 2018;160.
[7] Aghaali H, Ångstr
om H-E. A review of turbocompounding as a waste heat
recovery system for internal combustion engines. Renew Sustain Energy Rev
2015;49.
[8] W. Bou Nader, J. Chamoun, and C. Dumand, Thermoacoustic engine as waste
heat recovery system on extended range hybrid electric vehicles,Energy
Convers Manag, vol. 215, 2020.
[9] Lan S, Smith A, Stobart R, Chen R. Feasibility study on a vehicular thermo-
electric generator for both waste heat recovery and engine oil warm-up. Appl
Energy 2019;242.
[10] Di Battista D, Fatigati F, Carapellucci R, Cipollone R. Inverted Brayton Cycle for
waste heat recovery in reciprocating internal combustion engines. Appl En-
ergy 2019;253.
[11] Güven M, Bedir H, Anlas G. Optimization and application of Stirling engine for
waste heat recovery from a heavy-duty truck engine. Energy Convers Manag
2019;180.
[12] Trabucchi S, De Servi C, Casella F, Colonna P. Design, modelling, and control of
a waste heat recovery unit for heavy-duty truck engines. Energy Procedia
2017;129.
[13] Xu B, Rathod D, Yebi A, Filipi Z, Onori S, Hoffman M. A comprehensive review
of organic rankine cycle waste heat recovery systems in heavy-duty diesel
engine applications. Renew Sustain Energy Rev 2019;107.
[14] Feru E, Willems F, de Jager B, Steinbuch M. Modeling and control of a parallel
waste heat recovery system for euro-VI heavy-duty diesel engines. Energies
2014;7(10).
[15] Mahmoudi A, Fazli M, Morad MR. A recent review of waste heat recovery by
Organic Rankine Cycle. Appl Therm Eng 2018;143.
[16] Rijpkema J, Andersson S, Munch K. Thermodynamic cycle and working uid
selection for waste heat recovery in a heavy duty diesel engine. SAE Technical
Paper 2018-01-1371 2018.
[17] Preißinger M, Schw
obel JA, Klamt A, Brüggemann D. Multi-criteria evaluation
of several million working uids for waste heat recovery by means of Organic
Rankine Cycle in passenger cars and heavy-duty trucks. Applied Energy; 2017.
[18] Schilling J, Eichler K, K
olsch B, Pischinger S, Bardow A. Integrated design of
working uid and organic Rankine cycle utilizing transient exhaust gases of
heavy-duty vehicles. Appl Energy 2019;255.
[19] Li X, Song J, Yu G, Liang Y, Tian H, Shu G, Markides CN. Organic Rankine cycle
systems for engine waste-heat recovery: heat exchanger design in space-
constrained applications. Energy Convers Manag 2019;199.
[20] B. Xu, D. Rathod, A. Yebi, S. Onori, Z. Filipi, and M. Hoffman, A comparative
analysis of dynamic evaporator models for organic Rankine cycle waste heat
recovery systems,Appl Therm Eng, vol. 165, 2020.
[21] Alshammari F, Karvountzis-Kontakiotis A, Pesyridis A, Usman M. Expander
technologies for automotive engine organic rankine cycle applications. En-
ergies 2018;11.
[22] Guillaume L, Lemort V. Comparison of different ORC typologies for heavy-duty
trucks by means of a thermo-economic optimization. Energy 2019;182.
[23] Imran M, Haglind F, Lemort V, Meroni A. Optimization of organic rankine cycle
power systems for waste heat recovery on heavy-duty vehicles considering
the performance, cost, mass and volume of the system. Energy 2019;180.
[24] Galuppo F, Nadri M, Dufour P, Reiche T, Lemort V. Evaluation of a coupled
organic rankine cycle mild hybrid architecture for long-haul heavy-duty truck.
IFAC-PapersOnLine 2019;52.
[25] Huster WR, Vaupel Y, Mhamdi A, Mitsos A. Validated dynamic model of an
organic Rankine cycle (ORC) for waste heat recovery in a diesel truck. Energy
2018;151.
[26] Koppauer H, Kemmetmüller W, Kugi A. Modeling and optimal steady-state
operating points of an ORC waste heat recovery system for diesel engines.
Appl Energy 2017;206.
[27] Xu B, Rathod D, Kulkarni S, Yebi A, Filipi Z, Onori S, Hoffman M. Transient
dynamic modeling and validation of an organic Rankine cycle waste heat
recovery system for heavy duty diesel engine applications. Appl Energy
2017;205.
[28] NeunteuK, Stevenson PM, Hülser H, Theissl H. Better fuel consumption by
waste heat recovery. MTZ worldwide 2012;73(12).
[29] Freymann R, Strobl W, Obieglo A. The turbosteamer: a system introducing the
principle of cogeneration in automotive applications. MTZ worldwide
2008;69(5).
[30] Freymann R, Ringler J, Seifert M, Horst T. The second generation turbosteamer.
MTZ worldwide feb 2012;73.
[31] Seher D, Lengenfelder T, Gerhardt J, Eisenmenger N, Hackner M, Krinn I.
Waste heat recovery for commercial vehicles with a rankine process,21 st
Aachen Colloquium Automobile and engine technology. 2012.
[32] Koeberlein D. Cummins SuperTruck Program. Technology and system Level
Demonstration of highly efcient and Clean, Diesel powered Class 8 trucks.
Department of Energy Annual Merit Review; 2014.
[33] Dickson J, Damon K. Cummins/peterbilt Supertruck II. Department of Energy
Annual Merit Review; 2019.
[34] Allain M, Atherton D, Gruden I, Singh S, Sisken K. Daimler's super truck
program; 50 % brake thermal efciency. In: 2012 Directions in engine-
efciency and emissions research (DEER) Conference; 2012.
[35] Endo T, Kawajiri S, Kojima Y, Takahashi K, Baba T, Ibaraki S, Takahashi T,
Shinohara M. Study on maximizing exergy in automotive engines. SAE Trans:
Journal of Engines 2007;116.
[36] Marlok H, Bucher M, Ferrand N. Further Development of exhaust waste heat
recovery, vol. 13. ATZheavy duty worldwide; 2020. 3.
[37] Thantla S, Fridh J, Erlandsson A, Aspfors J. Performance Analysis of volumetric
expanders in heavy-duty truck waste heat recovery. SAE Technical Paper
2019-01-2266; 2019.
[38] Carstensen A, Horn A, Klammer J, Gockel J. Waste heat recovery in Passenger
Cars and trucks, vol. 80. MTZ worldwide; 2019. 4.
[39] Ekstr
om F. A mild hybrid SIDI turbo passenger car engine with Rankine waste
heat recovery. SAE Technical Paper 2019-24-0194; 2019.
[40] Rijpkema J, Ekstr
om F, Munch K, Andersson SB. Experimental results of a
waste heat recovery system with ethanol using the exhaust gases of a light-
duty engine. In: Proceedings of the 5th International Seminar on ORC po-
wer systems; 2019.
[41] Bettoja F, Perosino A, Lemort V, Guillaume L, Reiche T, Wagner T. NoWaste:
waste heat Re-use for greener truck. Transportation Research Procedia
2016;14.
[42] Furukawa T, Nakamura M, Machida K, Shimokawa K. A study of the rankine
cycle generating system for heavy duty HV trucks. apr 2014.
[43] Yang K, Zhang H, Song S, Zhang J, Wu Y, Zhang Y, Wang H, Chang Y, Bei C.
Performance analysis of the vehicle diesel engine-ORC combined system
based on a screw expander. Energies 2014;7.
[44] Zhang Y-Q, Wu Y-T, Xia G-D, Ma C-F, Ji W-N, Liu S-W, Yang K, Yang F-B.
Development and experimental study on organic Rankine cycle system with
single-screw expander for waste heat recovery from exhaust of diesel engine.
Energy 2014;77.
[45] Latz G, Erlandsson O, Skåre T, Contet A, Andersson S, Munch K. Performance
analysis of a reciprocating piston expander and a plate type exhaust gas
recirculation boiler in a water-based Rankine cycle for heat recovery from a
heavy duty diesel engine. Energies 2016;9(7).
[46] Shu G, Yu G, Tian H, Wei H, Liang X, Huang Z. Multi-approach evaluations of a
cascade-Organic Rankine Cycle (C-ORC) system driven by diesel engine waste
heat: Part A ethermodynamic evaluations. Energy Convers Manag 2016;108.
[47] Yu G, Shu G, Tian H, Huo Y, Zhu W. Experimental investigations on a cascaded
steam-/organic-Rankine-cycle (RC/ORC) system for waste heat recovery
J. Rijpkema, O. Erlandsson, S.B. Andersson et al. Energy 238 (2022) 121698
16
(WHR) from diesel engine. Energy Convers Manag 2016;129.
[48] Shi L, Shu G, Tian H, Huang G, Chen T, Li X, Li D. Experimental comparison
between four CO 2 -based transcritical Rankine cycle (CTRC) systems for en-
gine waste heat recovery. Energy Convers Manag 2017;150.
[49] Guillaume L, Legros A, Desideri A, Lemort V. Performance of a radial-inow
turbine integrated in an ORC system and designed for a WHR on truck
application: an experimental comparison between R245fa and R1233zd. Appl
Energy 2017;186.
[50] Alshammari F, Pesyridis A, Karvountzis-Kontakiotis A, Franchetti B,
Pesmazoglou Y. Experimental study of a small scale organic Rankine cycle
waste heat recovery system for a heavy duty diesel engine with focus on the
radial inow turbine expander performance. Appl Energy 2018;215.
[51] Alshammari F, Pesyridis A. Experimental study of organic Rankine cycle sys-
tem and expander performance for heavy-duty diesel engine. Energy Convers
Manag 2019;199.
[52] Shu G, Zhao M, Tian H, Huo Y, Zhu W. Experimental comparison of R123 and
R245fa as working uids for waste heat recovery from heavy-duty diesel
engine. Energy 2016;115.
[53] Taylor J. An introduction to error Analysis: the study of Uncertainties in
physical measurements. University Science Books; 1997.
[54] DieselNet: Emission test cycles - European Stationary Cycle (ESC).https://
www.dieselnet.com/standards/cycles/esc.php. Accessed: 2020-07-02.
[55] Danfoss. Nessie high pressure pumps for technical water, type PAH. Data
Sheet; 2007.
[56] Webb R, Nae-Hyun K. Principles of Enhanced heat transfer. CRC Press; 2005.
[57] Lemort V. Contribution to the Characterization of Scroll Machines in
Compressor and expander Modes. PhD thesis; 2008.
[58] Rijpkema J, Andersson S, Munch K, Thantla S, Fridh J. Experimental investi-
gation and modeling of a reciprocating piston expander for waste heat re-
covery from a truck engine. Applied Thermal Engineering; 2020.
[59] MathWorks. MATLAB, Version R2019a. 2019.
[60] Bell IH, Wronski J, Quoilin S, Lemort V. Pure and pseudo-pure uid thermo-
physical property evaluation and the open-source thermophysical property
library CoolProp. Ind Eng Chem Res 2014;53(6).
[61] Galuppo F, Reiche T, Lemort V, Dufour P, Nadri M. Organic Rankine Cycle
based waste heat recovery modeling and control of the low pressure side
using direct condensation and dedicated fans. Energy; 2020.
[62] Rijpkema J, Munch K, Andersson SB. Combining low- and high-temperature
heat sources in a heavy duty diesel engine for maximum waste heat recov-
ery using rankine and ash cycles. In: Junior C, Dingel O, editors. . Energy and
thermal management, air-Conditioning, and waste heat Utilization. Springer
International Publishing; 2019.
[63] Bredel E, Nickl J, Bartosch S. Waste heat recovery in drive systems of Today
and Tomorrow, vol. 72. MTZ worldwide; 2011.
[64] S. Quoilin, R. Aumann, A. Grill, A. Schuster, V. Lemort, and H. Spliethoff,
Dynamic modeling and optimal control strategy of waste heat recovery
Organic Rankine Cycles,Appl Energy, vol. 88, 2011.
[65] Grelet V, Reiche T, Lemort V, Nadri M, Dufour P. Transient performance
evaluation of waste heat recovery rankine cycle based system for heavy duty
trucks. Appl Energy 2016;165.
J. Rijpkema, O. Erlandsson, S.B. Andersson et al. Energy 238 (2022) 121698
17
... Rijpkema et al. [59] analyzed solutions for waste heat recovery (WHR) using an ORC to enhance engine efficiency, reducing CO 2 emissions in heavy-duty transport. Figure 13a shows the thermodynamic circuit related to a Rankine cycle, using water to recover waste heat from the diesel engine's exhaust. ...
... TEGs convert heat directly into electricity, offering a compact and maintenance-free solution, but their efficiency remains relatively low, typically between 1% and 3% (Zhao et al. [87], Burnete et al. [67]). ORC systems, by contrast, can achieve much higher recovery ratesoften between 3% and 10%-making them more suitable for heavy-duty applications (Rijpkema et al. [59], Di Battista et al. [62]). Some studies suggest that advanced hightemperature cycles, such as the Kalina cycle, could push efficiency up to 40% under optimal conditions (Roeinfard et al. [66]). ...
... (a) Scheme of the experimental setup, (b) equipment, and (c,d) exhaust evaporator presented in[59]. (a) Scheme of the experimental setup, (b) equipment, and (c,d) exhaust evaporator presented in[59]. ...
Article
Full-text available
Energy harvesting in the automotive sector is a rapidly growing field aimed at improving vehicle efficiency and sustainability by recovering wasted energy. Various technologies have been developed to convert mechanical, thermal, and environmental energy into electrical power, reducing dependency on traditional energy sources. This manuscript provides a comprehensive review of energy harvesting applications/methodologies, aiming to trace the research lines and future developments. This work identifies the main categories of harvesting solutions, namely mechanical, thermal, and hybrid/environmental solar–wind systems; each section includes a detailed review of the technical and scientific state of the art and a comparative analysis with detailed tables, allowing the state of the art to be mapped for identification of the strengths of each solution, as well as the challenges and future developments needed to enhance the technological level. These improvements focus on energy conversion efficiency, material innovation, vehicle integration, energy savings, and environmental sustainability. The mechanical harvesting section focuses on energy recovery from vehicle vibrations, with emphasis on regenerative suspensions and piezoelectric-based solutions. Specifically, solutions applied to suspensions with electric generators can achieve power outputs of around 1 kW, while piezoelectric-based suspension systems can generate up to tens of watts. The thermal harvesting section, instead, explores methods for converting waste heat from an internal combustion engine (ICE) into electrical power, including thermoelectric generators (TEGs) and organic Rankine cycle systems (ORC). Notably, ICEs with TEGs can recover above 1 kW of power, while ICE-based ORC systems can generate tens of watts. On the other hand, TEGs integrated into braking systems can harvest a few watts of power. Then, hybrid solutions are discussed, focusing on integrated mechanical and thermal energy recovery systems, as well as solar and wind energy harvesting. Hybrid solutions can achieve power outputs above 1 kW, with the main contribution from TEGs (≈1 kW), compared to piezoelectric systems (hundreds of W). Lastly, a section on commercial solutions highlights how current scientific research meets the automotive sector’s needs, providing significant insights for future development. For these reasons, the research results aim to be guidelines for a better understanding of where future studies should focus to improve the technological level and efficiency of energy harvesting solutions in the automotive sector.
... One of these concerns relates to the recovery of waste heat discharged from the engine. The main technologies available for waste heat recovery are as follows: thermoelectric generators [2]; thermoacoustic generators [3]; turbocompound systems [4]; technologies using thermal cycles (steam Rankine [5]; organic Rankine [6]; Stirling [7] and Brayton with air [8] or with CO 2 [9]) and refrigeration systems [10]). The purpose of waste heat recovery systems is to recover as much heat as possible to convert it into electricity or cold. ...
... This paper addresses the recovery of waste heat from the exhaust gas of the internal combustion engine of a series of hybrid electric buses using the split-flow sCO2 recompression Brayton cycle. A mathematical model of the heat exchange 6 The cyclopentane ORC cycle follows the split-flow sCO 2 recompression Brayton cycle in terms of net power, thermal efficiency, and heat recovery efficiency. The higher performance compared to the steam Rankine cycle was due to the internal heat recovery. ...
Article
Full-text available
Waste heat recovery from exhaust gas is one of the most convenient methods to save energy in internal combustion engine-driven vehicles. This paper aims to investigate a reduction in waste heat from the exhaust gas of an internal combustion engine of a serial Diesel–electric hybrid bus by recovering part of the heat and converting it into useful power with the help of a split-flow supercritical CO2 (sCO2) recompression Brayton cycle. It can recover 17.01 kW of the total 33.47 kW of waste heat contained in exhaust gas from a 151 kW internal combustion engine. The thermal efficiency of the cycle is 38.51%, and the net power of the cycle is 6.55 kW. The variation in the sCO2 temperature at the shutdown of the internal combustion engine is analyzed, and a slow drop followed by a sudden and then a slow drop is observed. After 80 s from stopping the engine, the temperature drops by (23–33)% depending on the tube thickness of the recovery heat exchanger. The performances (net power, thermal efficiency, and waste heat recovery efficiency) of the split-flow sCO2 recompression Brayton cycle are clearly superior to those of the steam Rankine cycle and the organic Rankine cycle (ORC) with cyclopentane as a working fluid.
... [5][6][7] Heat pipes and thermoelectric generators as a combination have the capability to form solid, robust, passive WHR systems. 8,9 Exhaust gas recirculation (EGR) is another effective means to make use of waste heat and reduce NO x formation. 10 Chemical heat storage (CHS) and the organic rankine cycle assist in utilizing the waste heat. ...
... 10 Chemical heat storage (CHS) and the organic rankine cycle assist in utilizing the waste heat. 8,9 Experiments using cascaded TES for WHR from a diesel engine exhaust indicated that "the majority of heat is present at higher temperatures for discharge applications and virtually at the PCM phase change temperature." 10 Heat pipes with TES help enhance the TES unit's performance. ...
Article
Loop heat pipe (LHP) encased in phase change material (PCM) incorporated annular to catalytic converter (CC) is proposed to augment the performance of the “thermal energy storage” (TES). LHP are designed to extract surplus heat from the exhaust discharge, thereby reducing the amount of exhaust heat emitted into the atmosphere. A four-cylinder IC engine’s CC is considered with Mg 70 Zn 24.9 Al 5.1 , paraffin oil as PCM and working fluid for the heat pipe are evaluated. A comparative experimental investigation was performed considering two models for CC with and without heat pipe integrated in the TES for (i) distribution of average PCM temperature (ii) energy, exergy efficiency and distribution (iii) effectiveness (iv) instantaneous waste heat recovered during heat storage. Transient cycles at 2000 r/min with varying load conditions were run considering city drive conditions. Heat storage for CC with heat pipe encased in TES was found to be 35% faster, whereas discharge time for CC without heat pipe developed in TES was found to be longer. For melt fraction 1, maximum exergy efficiencies of 71% and 69% were observed for the TES unit with and without heat pipe. Heat pipes are observed to help extract a considerable amount of surplus heat from exhaust and reduce the exhaust gas outlet temperature released into the atmosphere by nearly 130 – 40°C under various load circumstances.
... Despite the significance of road cargo transportation for the country's development, this dependence on fossil fuel-generated energy contradicts global policies aimed at reducing emissions. Mahesh et al. [12] emphasize that trucks are the most polluting vehicles, accounting for 5% of the total greenhouse gas emissions in Europe [13]. ...
Article
Full-text available
This article presents an original research methodology that combines insights from patents and academic research, offering a unique perspective on energy recovery technologies for trucks equipped with refrigeration units. The purpose of the study is to perform a functional analysis of existing solutions and to suggest a mechanism for exposing unexplored areas and opportunities for innovation. To achieve this goal, a systematic opportunity scan is presented, investigating patents and conducting a state-of-the-art search of existing technologies. This scan classifies a diverse range of solutions, elucidating their interconnections and providing an overview of the existing technological area, covering system components and technical trends. Thus, the main functions and components are listed, as well as the system requirements. Once the functions have been surveyed, a morphological matrix is proposed, and five main functions are analyzed. This methodology makes it possible to list the majority of the possible solutions for the functions analyzed, taking into account the components observed in the literature review and patents, including new components raised by the research group. Finally, with the morphological matrix structure, it was possible to combine unexplored elements, achieving innovative solutions.
... Moreover, its integration in a vehicle results in constraints that limit the recovery: weight increase, backpressure effect on the engine, cold sinks, final energy storage, and utilization on board [31]. Generally, real working units demonstrate performances far from those theoretically predicted, and technological advancement in components and integration [32] is needed. Efforts to increase the power recovered from exhaust gases are welcomed because they contribute to reducing fuel consumption and CO 2 emissions. ...
Article
Energy recovery has become an important solution that drives the transition to more sustainable propulsion, since the amount of thermal energy available from exhaust gases is huge. This energy can be transformed into electrical energy (without significantly affecting the powertrain) and directly used, for instance, in hybrid propulsion or driving auxiliaries. A turbocharged diesel engine equipped with a variable geometry turbine (VGT) was tested to assess the maximum energy recoverable from exhaust gases through two different recovery stages. The first was achieved using the pressure difference between the value at the exhaust valves and the atmospheric datum (turbo-compounding). The second recovery stage was achieved thanks to the temperature of the exhaust gases after the first recovery, and performed using an organic Rankine cycle (ORC)-based power unit. Therefore, these two combined stages of energy recovery were experimentally investigated in the medium–low load region of the ESC-13 homologation test, which is representative for real driving of heavy-duty engines. The first stage was performed on the turbocharging system by recovering the energy lost inside the VGT, through an additional turbine that operates in parallel with the main turbine that drives the compressor. It facilitates the recovery of a mechanical power of up to 3 kW, which was approximately equal to 5 % of the engine brake power at a specific medium–low load. The second stage was performed with an ORC-based unit bottomed to the first recovery section, exploiting the fact that after the first recovery, exhaust gases still have a high temperature, which can be used to feed the additional ORC based recovery unit. This second stage adds up to 3.5 kW of recovered mechanical energy, which represents 5 % of the engine brake power at the same medium–low engine load operating points. Therefore, a total of 10 % of the engine power was recovered in the two stages, which are characterized by proven technologies. Considering the engine working point at maximum power, the value was also higher, the combined recovery achieved a mechanical power of approximately 14 % of the engine brake power. In this study, the detrimental effects related to the engine backpressure produced by the two recovery units and the additional weight of the vehicle were assessed, demonstrating a net overall specific fuel consumption reduction of approximately 5–7% in the medium–low operating region of the engine considered, and higher than 8% at the maximum engine power. The increase in complexity related to the two recovery stages invites to consider this technology for heavy-duty engines for long-hauling vehicles, in which the detrimental effects does not significantly affects fuel consumption.
Article
Full-text available
Diesel engines play a critical role in numerous sectors due to their robustness and efficiency but contribute significantly to environmental pollution through exhaust emissions. These emissions contain substantial thermal energy, which, if not harnessed, exacerbates air quality issues and global climate change. The Organic Rankine Cycle (ORC) offers a promising solution for recovering this waste heat and converting it into mechanical or electrical power. This study conducts thermodynamic simulations using MATLAB to analyze ORC systems for diesel engine exhaust heat recovery, focusing on thermal efficiency, power generation potential, and environmental impact of different refrigerants, specifically R-141b, R-245fa, and R-123. Results indicate that ORC performance is significantly influenced by the refrigerant’s critical temperature. R-141b demonstrates the highest thermal efficiency at 13.13% with a heat source temperature of 120°C but also has the highest environmental impact, contributing 3,090 kgCO2 eq/year. In contrast, R-123 shows the lowest environmental impact at 231 kgCO2 eq/year for direct TEWI and 7.73 kgCO2 eq/year for indirect TEWI, with a slightly lower thermal efficiency of 12.73%. R-245fa, with the lowest efficiency at 12.14%, also has a substantial environmental impact. This research provides insights into optimizing energy recovery from diesel engine waste heat and advancing sustainable energy solutions.
Article
Industrial Carbon black production is a petrochemical process that has high level of waste heat and emissions. Main points of CB process include furnace and stack lines. Also, the use of different feedstocks can be effective in level of waste heat and amount of emission pollutants. So the main objective of this research use from the pinch technology in CB process in order to minimize the level of waste heat by utilization of waste heat recovery (WHR) applications for power generation. By numerical investigation and modelling of WHR in furnace and stack points, can be defined best location for power generation in CB process. Moreover, through the analysis of Organic Rankine Cycle (ORC) systems integrated with WHR from both energy and exergy perspectives, second target to enhance their thermal efficiency. The results indicate that the presence of blowers on stacks leads to an increase in RoHR and RoHC. In order to determine the optimal location for power generation cycle installation, an ORC energy and exergy analysis was conducted. The results showed that the energy and exergy values at the SLC stack were higher than those at the HLC point.
Article
The Goswami combined power/cooling cycle is an appropriate way to convert energy efficiently. However, in the conventional Goswami cycle (CGC), the refrigeration is not available when the absorber temperature is high. In the present study, a mechanically driven compressor is used to compress the vapors leaving the turbine. This design creates a line independent of the absorber temperature. An idealized cycle approach was used with the assumptions adopted by Goswami. As a result, under the same typical operating conditions, the thermal efficiency of the novel cycle is 27.52% with a net useful effect of 135.2 kW, while the thermal efficiency of CGC is 23.54% with a net useful effect of 99.47 kW. Moreover, at equal total heat input, the working fluid mass flowrate in the novel cycle is less than the CGC. Ultimately, the refrigeration limits highlighted in the CGC were overcome in the novel cycle which allows the availability of refrigeration regardless of the increase in absorber temperature. This advantage, combined with the improved thermal efficiency, suggests a short-to-medium-term payback time of the additional investment cost of the added components and a long-term profitability of the system.
Article
Full-text available
Waste heat recovery using an (organic) Rankine cycle has the capacity to significantly increase the efficiency of heavy-duty engines and thereby reduce fuel consumption and CO2 emissions. This paper evaluates a reciprocating piston expander used in a Rankine cycle for truck waste heat recovery by quantifying its performance on the basis of experimental results and simulations. The experimental results were obtained using a setup consisting of a 12.8 L heavy-duty Diesel engine connected to a Rankine cycle with water and are used to calibrate a semi-empirical expander model. At an engine power between 75 and 151 kW, this system recovered between 0.1 and 3 kW, resulting in an expander filling factor between 0.5 and 2.5, and a shaft isentropic effectiveness between 0.05 and 0.5. The calibrated model indicated that the heat loss (16 %), mechanical loss (6 – 25 %), pressure drop (13 – 42 %), and leakage (25 – 75 %) all contributed significantly to the expander performance loss. A simulation study with acetone, cyclopentane, ethanol, methanol, and R1233zd(E), showed that a change of working fluid significantly impacts the expander performance, with the filling factor varying between 0.5 and 2.2 and the effectiveness between 0.01 and 0.5, depending on the working fluid, expander speed, and pressure ratio. The results of the optimization of the built-in volume ratio and inlet valve timing during a typical long haul driving cycle showed that acetone and R1233zd(E) provided the highest available power around 3 kW absolute, or 2.2 % relative to the engine. The main contributions of this paper are the presentation of experimental results of an engine coupled to a Rankine cycle, and the quantification of performance losses and the effect of working fluid variation using an adapted semi-empirical expander model, which allows for a selection of the working fluid and geometrical modifications giving optimal performance during a long haul driving cycle.
Conference Paper
Full-text available
Organic Rankine cycle (ORC) waste heat recovery (WHR) systems have the potential to improve the efficiency of modern light-duty engines, especially at highway driving conditions. This paper presents and discusses the experimental results of an engine connected to a compact ORC-WHR system with ethanol, suitable for integration in a modern passenger car. The aim is to show the added value of this ORC-WHR system for passenger cars by presenting the experimental results with the focus on the expander power output. The experimental setup consists of a Volvo Cars VEP-4 gasoline engine, which has an evaporator integrated in the exhaust pipe. During operation, one of two different states can be selected: electrical feedback (EFB) or mechanical feedback (MFB), where the expander can be either coupled to a 48V generator (EFB) or directly to the engine (MFB). Control strategies were developed to allow for operation of the system without interference of the driver. The results show that the current setup and control strategies can be successfully employed with significant expander power outputs for both MFB and EFB. The expander power outputs, similar for both states, go up to 2.5 kW, recovering 6.5 % of the available exhaust energy and giving more than 5 % improvement in fuel consumption.
Article
Full-text available
Effective recovery of heavy-duty vehicle waste heat is a key solution toward meeting the increasingly stringent fuel economy and CO2 emission standards. Different from previous publications, this paper presents a preliminary introduction of organic Rankine cycle waste heat recovery (ORC-WHR) in heavy-duty diesel (HDD) vehicle applications in the past decade. It presents a wide range of topics in the HDD vehicle ORC-WHR system development , including system architecture evaluation, heat exchanger selection, expander selection, working fluid selection, power optimization, control strategy evaluation, simulation and experimental work overview, and limiting factors. In the system architecture selection, the tradeoff between fuel savings and system complexity dominates. In the heat exchanger design, besides the heat exchanger efficiency, transient evaporator response is critical factor for the system control and performance. The expander type and configuration is closely coupled to the expander power output type (i.e. electricity or mechanical power). WHR power production is most sensitive to working fluid mass flow rate, with less sensitivities to expander speed and condenser coolant mass flow rate. The integration of ORC-WHR control with engine control shows potential to improve the waste heat recovery system performance. The simulation studies predict higher power recovery levels than that in experimental work (0-60kW vs. 0-14kW), which could result from the large number of heat resources, optimistic expander and pump efficiencies and neglected heat losses in the simulations.
Article
Organic Rankine cycle based waste heat recovery has been studied over the past decade as a potential solution to reduce fuel consumption and fulfill the requirements of upcoming regulations on CO2 emissions for heavy-duty trucks. This study, focusing on a particular configuration of the system, using direct condensation and dedicated fans, presents the models of all the components in the system and validation of the evaporator and fan model according to experimental results. A special attention has been paid to the effectiveness of the fan speed and condensation pressure control to increase the net power output of the direct condensed ORC. Road cycle simulations are run over an internal Volvo cycle that is suitable for the present application, using a "base" control strategy; afterwards a sensitivity analysis and optimization, involving the low pressure loop of the system, are performed, leading to better performance in terms of net power produced by the Rankine system.
Article
Waste heat recovery (WHR) systems suggest a promising solution for reducing vehicle CO2 emissions in order to meet the CAFE targets by 2025. This paper presents a methodology to improve the overall efficiency of a combined cycle machine consisting of a reciprocating internal combustion engine (ICE) coupled to a thermoacoustic (TAE) machine used for thermal-to-electric WHR. It investigates the potential of calibrating the ICE at some specific points of its engine map in order to achieve an optimal overall efficiency when coupled to the bottoming thermoacoustic cycle. A three-cylinder gasoline engine is modeled using GT suite code and the effect of spark timing delay on exhaust temperature, exhaust flow and engine brake efficiency are compared to real engine test bench values. Three bottoming thermoacoustic configurations coupled to this ICE are modeled and calibrated according to test results performed on a thermoacoustic machine prototype. The resulting electrical power recovery from the exhaust gas is analyzed and assessed. A Range Extender Hybrid Electric Vehicle (EREV) is considered and fuel consumption is simulated on the WLTC. The results revealed an added value for adding a multi-module TAE in series, to optimize heat recovery with a potential of consumption reduction up to 7.6%. Results have also shown interest in delaying the ignition, enabling higher exhaust temperature and mass flow rate which tend to positively impact the TAE machine. The proposed method is also beneficial in the way that it avoids knocking problems and enables the future design of higher compression ratio engines for auxiliary power unit on EREV.